The helix dipole Jasper Minkels, 0610437 Introduction Throughout the scientific literature on protein research of the last thirty years, the α-helix dipole is mentioned in a number of cases where its partial charge on both the C-terminal and the N-terminal ends is thought to be involved in various biological processes on a molecular level. In order to understand the cause of this dipole moment, one must look at the structure and the geometry of the α-helix. The α-helix is a right handed coiled structure. Each amino acid makes a turn of about 100 degrees and hence it requires 3.6 residues to make a full turn. Amino acids that wind around the axis form hydrogen bounds with each other. The N-H group of each amino acid forms a hydrogen bond with the C=O of the amino acid that is located four places earlier in the helix. The vast number of hydrogen bonds that is formed is actually the underlying reason for the forming of this secondary protein structure. Although the principles behind the occurrence of the helix dipole have been described earlier, its role in protein structure and function was first reviewed by Hol in 1985. He states that the helix dipole originates in the dipole of the individual peptide unit. The charge distribution within such a unit is pictured in Figure 1. Its direction is parallel to the N-H and C=O bonds. It has been shown that in an α-helix around Figure 1: Peptide charge distribution, 97% of all peptide dipole moments point in the direction of the numbers in units of the elementary helix axis, the dipole is therefore quite insensitive to the φ and ψ charge[8] angles. The C=O groups are in a slightly upward direction (toward the C-terminus) and the N-H groups are in a downward direction (toward the N-terminus) and this gives rise to a small dipole. The aggregate effect of all individual dipoles in an α-helix is a negative dipole moment at the C-terminal end, and a positive dipole moment at the N-terminal end of the helix. The dipole moment of an individual peptide unit is about 3,46 D which equals 0,72 e Å or 0,5 e per 1,5 Å. Since the axial shift per residue in an α-helix is also 1,5 Å, all dipoles cancel out except for the C- and N-terminal ones. However, a dipole moment is also affected by the local environment and especially hydrogen bonds are known to increase the value of dipole moments. Both calculations and experimental measurements confirm that the dipole moment is indeed increased, by 25 to 50%. Although the amount of known protein structures and biological mechanisms has enormously expanded since then, Hol already distinguished a few roles for the helix dipole in proteins that still prove to be relevant. In his review he discusses the role of the helix dipole in glutathione peroxidase, rhodanese, subtilisin, triose phosphate isomerase, glyceraldehyde-3phosphate dehydrogenase, p-hydroxybenzoate hydroxylase, glutathione reductase, thioredoxin and papain. In these proteins, the helix dipole is used for binding stabilizing (intermediary) charges varying from anions to active site residues. Hol also analyzed twenty proteins that were known to bind phosphate groups. The majority of these proteins combined the N-terminal dipole and positively charged residues for this, although a few proteins such as flavodoxin depended solely on a helix dipole and neighbouring N-H groups for phosphate binding. It was found that the phosphate binding helices were often part of a βαβ-motif. However, the individual amino acid sequences mostly differed, except for a highly conserved glycine that was incorporated to provide space for the approaching phosphate moiety. Hol thus concluded that the electrostatic interactions, and not a specific sequence, was essential to the phosphate binding mechanism. Further research aimed to analyze if the binding by these anions was due to the possibility of forming large amounts of hydrogen bonds concluded this was not the case, thereby confirming his conclusion. At the time no proteins were known that applied a similar mechanism to DNA or RNA binding. Hol further concluded in his review that the α-helix dipole may explain the most frequently observed folding patterns in proteins. He also included the dipole of parallel β-strands in his model, but further research proved this dipole to be non-existent. His conclusions in summary: (i) In all-helical proteins, helices tend to be anti-parallel because of a favourable interaction between alpha-helix dipoles. This also explains the frequent occurrence of alpha-alpha-units; (ii) In αβ-proteins, a favourable interaction occurs between the β -dipoles in the center of the molecule and the α-helix dipoles arranged in an anti-parallel way around the parallel beta-strands. Such an interaction would also explain the stability of the βαβ -unit; (iii) In all-β proteins the favourable αβ-dipole interactions are absent and β-strands tend to be anti-parallel as an unfavourable interaction between parallel β-dipoles would otherwise occur. Finally, Hol performed calculations on the interaction between two α-helices in vacuum and concluded that the most important parameters are the helix length, the distance between the two helices and the Ω-angle. These parameters are pictured in Figure 2. Hol recognized four different related environments: bulk water, bulk protein, the outer protein surface and the proteins first hydration shell. Due to different mobilities and dielectric properties Hol rightfully concluded that exact calculations on interactions in proteins aren’t straightforward as no dielectric constant can be applied. We will see whether Hol was right in stating that “the α-helix dipole plays an important role in modulating the properties of Figure 2: Visualization of the parameters used by Hol[8] several enzymes and in defining the mode of coenzyme [8] binding by numerous proteins”. Channels Ion channels The electrostatic effects of the helix dipole are often described as relevant in multiple classes of ion channels. Doyle et al. solved the protein structure of the potassium channel from Streptomyces lividans at a 3,2 Å resolution. Its amino acid sequence is similar to that of other potassium channels, both voltage gated and cyclic nucleotide gated, in various organisms. Molecular cloning and mutagenesis experiments have reinforced the conclusion that all K+ channels have essentially the same pore constitution, sharing a critical amino acid sequence that is vital for distinguishing K+ from Na+ ions. This specificity and the high conductivity of the channel, up to 108 ions per second, are the main points of interest of the authors. The following information can be extracted from their article[4]. The potassium channel is a tetramer consisting of four identical units with two membrane-spanning regions each. Of these two, one transmembrane helix faces the central pore while the other faces the lipid membrane. The inner helices are tilted with respect to the membrane normal by about 25 Figure 3: Ribbon representation of the degrees and are slightly kinked. The inner helices of all four tetramer as an integral membrane protein. [4] subunits pack against each other as a bundle near the intracellular aspect of the membrane, giving the appearance of what the authors call an “inverted teepee”[4]. Figure 3 is a schematic representation of the channel. Negatively charged acidic amino acids are located near both entryways of the channel. This ensures a high concentration of cations. From inside the cell the pore begins as a tunnel and then opens into a wide cavity near the middle of the membrane. Up to this part a passing K+ ion can remain largely hydrated. Along with these water molecules the helix dipoles of the four pore helices, with their carboxyl end pointed towards the center of the cavity, help to lower the energy barrier although they are 8 Å away from the cavity center, a point to which we will return later on. Figure 4 depicts the cavity and the surrounding helices. Whereas the selectivity filter is lined with polar amino acids, the pore is lined with hydrophobic groups. The authors propose that the lack of possible interactions between these groups and the passing ion improves the throughput rate. Another improvement follows from the structure of the selectivity filter. It consists of two binding spots. While one ion is bound quite tight, the arrival of the second ion on the nearby binding spot repulses the first one, thereby pushing it through the selectivity filter. A molecular spring that ensures the widening of the selectivity filter accounts for the fact that a dehydrated Na+ ion, which has a 0,95 Å radius compared to 1.33 Å for K+, cannot be Figure 4: Both the aqueous cavity and the helices (four in total) stabilize the ion (green) properly stabilized. Hence the fact that the channel is over in the K+ channel.[4] 10.000 times more permeant to potassium ions.[4] Roux and MacKinnon quantified the stabilization energies in a similar type of potassium channel, the KscA K+ channel, using the Born theory of solvation[2]. The free energy for transferring an ion from bulk water into a cavity of radius R and dielectric constant ε w embedded in a medium of low dielectric constant ε m is given by: ΔG = (1 / 2) (Q2 / R) (1 / ε m -1 / ε w)[19] The transfer of a potassium ion from an aqueous solution to a water filled sphere (radius 5 Å, ε w = 80) surrounded by hydrocarbons (ε m = 2, which corresponds to a 25 Å membrane) gives a value of 16.2 kcal/mol, which is a decrease of over 40 kcal/mol compared to a transfer to a hydrocarbon environment. When the actual shape of the cavity and the fact that the bulk water is not infinitely far away is taken into account the authors arrive at a value of 6.3 kcal/mol. The fact that two K+ ions are present in the selectivity filter increases this value with another 10 kcal/mol. The transfer energy drops to –8.5 kcal/mol when all the atomic charges of the protein and the presence of the selectivity filter ions are turned on. The authors stated that dielectric shielding by water molecules in the cavity minimizes the electrostatic influence of the helix dipole moments, 8 Å away from the cavity center. They add, however, that the low dielectric membrane environment contributes significantly to the amplification of cation stabilization. They calculated that the helices are responsible for eighty percent of the stabilization, leaving the other twenty percent to a Thr carbonyl group.[19] Using the model proposed by Doyle et al., Zhou et al. sought to answer two additional questions regarding the Ksca K+ channel. The first is how the K+ channel mediates the transfer of a K+ ion from its hydrated state in solution to its dehydrated state in the selectivity filter. This is relevant to all mechanisms of selective ion transport. Because dehydration in the wrong environment is energetically costly, the authors went to look for a set of mechanisms designed to handle hydrated K+ ions and also mediate their dehydration. The second question addressed in this study is related to the cellular environment in which K+ channels operate: inside the cell the K+ concentration is greater than 100 mM, whereas on the outside the K+ concentration is usually less than 5 mM. By solving structures at both high and low K+ concentrations, the authors were able to conclude that two conformations of the channel occur[23]. To return to the first question; The hydration of K+ in the cavity and the conformation of the carbonyl atoms arranged to replace the hydration appear to be precisely regulated. The cavity in KscA and most other K+ channels is lined mainly by hydrophobic amino acids from the inner helix and hence no strong hydrogen-bonding donor or acceptor groups are available for the water molecules surrounding the ion. Consequently, water in the cavity is available to interact strongly with the K+ ion. Eight water molecules surround the K+ ion in a square antiprism conformation. This orientation will later be mimicked in the selectivity filter by carbonyl groups. Regarding the question of how this precise ordering is conducted, the authors conclude that inspection of the cavity wall shows that the order is probably imposed by a sum of very weak, indirect hydrogen-bonding interactions mediated by certain chemical groups, and perhaps the carbonyl oxygen atoms from the pore and inner helices. The cavity achieves a very high effective K+ concentration (2 M) at the center of the membrane, with the K+ ion positioned on the pore axis, ready to enter the selectivity filter.[23] Dutzler et al. Determined the structure of two CIC chloride channels. The resolution of these structures, originating from Salmonella enterica serovar typhimurium and Escherichia coli, is 3.0 and 3.5 Å respectively. While the biological role of this class of chloride channels remains unknown in many cases, they are widespread in both eukaryotes and prokaryotes. In vertebrates they account for stabilizing the membrane potential in skeletal muscle cells and regulation of excitability. Because of the sequence conservation that is observed among different organisms it is highly plausible that the mechanisms of ion submission are equal. Both channels are double-barrels, consisting of two dimers bound in an antiparallel fashion. This antiparallel arrangement differs from the architecture of K+ channels, but is also found in aquaporins. This architecture enables both ion stabilization by N-termini at the inside of the channel and location of C-termini in the aqueous environment outside of the membrane, leading to both ion stabilization and avoidance of the energetically costly burial of the C-terminus in an hydrophobic environment. Although at present only dimers appear to be functional channels each unit has its own pore and selectivity filter.[5] A comparison of amino acid sequences of both units show that the units are slightly correlated especially with respect to glycines. Transmembrane helices of both units wrap around a common center, thereby bringing the ends of different helices together. The Ntermini of two helices appear to be directed towards a Cl- binding spot. Other atoms thought to be involved in ion coordination are Ile and Phe main-chain amide nitrogen atoms Figure 5: The anti-parallel orientation allows like helix ends to point at the and sidechain oxygen atoms from selectivity filter from opposite directions [5] Ser and Tyr. The authors suggest that coordination by partial charges permits rapid diffusion rates, whereas using full charges would result in a deep energy well. A second ion binding spot is located at the extracellular side of the channel. This spot is occupied by a carboxylate ion and the authors suggest that this spot is important for the strong coupling between gating and ion conduction that is a general characteristic of CIC Cl- channels. Chloride ions compete for this second binding site. Upon binding, certain conformational changes take place in the selectivity filter so that the sidechain of a conserved Glu blocking the pore swings out of the way. The chloride bound at the inner site is thought to play an important role in the ion selectivity of the channel, thereby giving rise to properties such as voltagedependent gating of some ClC Cl- channels.[5] Faraldo-Gomez and Roux examined the contribution to the electrostatic binding free energy of local and non-local interactions of EcClC, a ClC-type chloride channel homologue from Escherichia coli and made a comparison with the stabilization in K+ channels and sulphate and phosphatebinding proteins (SBP and PBP). While the helix dipole is often mentioned as an important contributant to energetic stabilization of ion species, as we have already seen, various theoretical studies, suggest otherwise, concluding that the stabilization is due to local interactions. Hence the comparrison made by the authors. The authors conclude that the contribution of helix macrodipoles to chloride binding in EcClC is only marginal and is analogous to the binding mechanism of the sulphate and phosphate-binding proteins. The transmembrane potassium channel KcsA in contrast seems to profit from helical stabilization in a significant way. The following reasons may account for this. Firstly, the low dielectric environment provided by the membrane for example helps to magnify the stabilizing effect, a fact that was also recognized by Roux and MacKinnon[19]. Calculations based on a situation where the membrane is absent show a decrease in stabilization by 50%. The distant ionbackbone interactions with the cavity ion in K+-channels, with the helix being 8 Å from the ion, come mainly from the pore helices and are prominent simply because more proximal interactions do not occur. EcClC is topologically more similar to SBP and other soluble ion-binding proteins, whereas its environment is analogous to that of KcsA. Binding of anions within PBP, SPB and ClC channels is largely due to favorable electrostatic interactions with the backbone of the protein. These interactions must at least balance the energetic cost of desolvation. In both PBP and SPB the binding site is located between the N-termini of three α-helices. Calculations show that only the first (80%) and second (95% for the first two turns together) turn contribute to the stabilization. The different relative importance of helical macrodipoles for the energetic stabilization of ions in K+-channels and the SBP and PBP proteins can therefore be understood on the basis of the distinct architecture and environment of these proteins.[6] Chatelain et al. obtained results contradicting the results of these calculations, and hence the mechanistic models of Doyle and Dutzler[4, 5]. They conclude that helix dipoles have a negligible role in electrostatic stabilization and channel function. They introduced several mutations in yeast potassium channels. The C-termini of helices were modified by introducing positively charged residues at a conserved position, thereby canceling out the electrostatic effect of the helix. Diffusion rates vary for each mutation but generally mutated channels function well, indicating that potassium ions are largely blind to electrostatic perturbations at the C-terminal end of the pore helix. They also question the conclusion of Roux and MacKinnon[19], stating that electrostatic effects over a distance of 8 Å even exceed the distance covered by ion-dipole interactions. Chatelain et al. suggest that the stabilizing effects of the helix, if any, arises from unpaired hydrogen bonding groups near the termini that act over very short distances.[3] Aquaporins The usage of helix dipoles in selectivity filters is also proposed by Murata et al. Using electron crystallographic data, he and his coworkers determined the structure of the aquaporin AQP1 at a 3.8 Å resolution. This is a water-selective membrane pore found in red blood cells and renal proximal tubules. Other members of the aquaporin family are found throughout nature where they are involved in numerous physiological processes. The complete channel consists of a tetramer of identical subunits, each being 269 amino acids in length and forming three membrane-spanning helices. Besides its exceptional water permeability of over 3 x 109 molecules per second, its ability to prevent any protons from crossing is intruiging. It is known that water can readily cross cell membrane as a continuous unbroken column of molecules within an open pore. This hydrogenbonded chain of water molecules normally conducts protons with great efficiency. Based on the structure the authors conclude that two pore helices, HB and HE, play a vital role in this mechanism for distinguishing. Because of the resolution no water molecules (2.8 Å) could be distinguished, so this conclusion remains hypothetical. The shape of the AQP1 pore is the opposite of the pore shape formed by the potassium channel, having a constriction in the centre of the membrane and wide openings at the membrane surfaces. This constriction results in a high dielectric barrier that repels ions, while it allows for penetration by neutral solutes. Both pore helices have an adjacent Asn-ProAla motif that is highly conserved. Prolines of these motifs interact to ensure the orientation of the pore helices and to maintain the diameter of the pore at 3 Å. Due to the positive electrostatic field generated by the dipole moments of the pore helices, the oxygen atom of a passing water molecule orients to the side of this motif, whereas the hydrogen atoms are directed towards hydrophobic residues lacking the ability to form hydrogen bonds. This comes at an energetic cost of 3 kcal/mol, since only one hydrogen bond is lost compared to a bulk water environment. Because of this hydrogen bond isolation, the water column is disrupted. Thus, protons no longer can pass the channel. This mechanism is brightly illustrated in Figure 6[14]. Fu et al. state that at present, ten aquaporins of the human family are known, all of which are permeable to water. AQP3, AQP7, and AQP9 are also permeable to glycerol. Glycerol permeabilty results from a mchanism quite similar to that of AQP1, albeit that carbon atoms of glycerol are aligned to a hydrophobic wall whereas the oxygen atom of water is coordinated by a helix dipole. In eukaryotes many of the family members are regulated by phosphorylation, pH, osmolarity, or binding of other proteins or ligands. AQP6 conducts cations at pH values below 5,5.[7] Quantifying the dipole moment In 1988, Šali, Bycroft and Fersht quantified the electrostatic stabilization energy of various forms of barnase, a small ribonuclease found in Bacillus amyloliquefaciens, by NMR-titration. The enzyme contains a histidine at the C-terminal end of a twelve amino acid long α-helix. The authors reported that the pKa of the wild-type enzyme is Figure 6: Representation of AQP1 and its mechanism to 7.9, whereas titration of the urea-denatured disrupt the water column by forming alternative hydrogen enzyme results in a pK of only 6.3. At a neutral a bonds with the channel.[14] pH, the stabilization is therefore 2.1 kcal/mol. The authors also conducted an experiment adding 1M KCl, a solution known to counter the charges of individual protein sidechains, to rule out these effects in the stabilization. No sidechain involvement was found, as the structure already predicted.[20] Another effort to calculate the strength of the electrostatic effect of the α-helix dipole was carried out by Lockhart and Kim. They introduced a probe, 4-(methylamino)benzoic acid (MABA), at the amino terminus of a protein whose absortion bands could be measured. Because the difference between the ground state and the excited state of MABA is known, the electric field could be determined by measuring the shift in the absorption band using Stark effect spectroscopy.[11] The authors synthesized a range of peptides that differed in length, amino acid composition, resulting in folding and non-folding proteins and the occurrence of Arg. 2D-NMR and circular dichroism were applied to check that the proteins that should be were properly helical. The authors checked the position of the probe using NMR. It was found that in the presence of a helical structure, the absorption band of the MABA probe is shifted by -5 nm. This band shift corresponds to a change in the transition energy of 1.6 kcal/mol. Similar values were found with other proteins, indicating that the observed field is primarily produced by backbone dipoles, not sidechain charges, and that the magnitude of the field is independent of the helix length between 21 and 41 residues. Although diminished by the environment, the field measured at the boundary between the peptide and water is still an order of magnitude stronger than expected based on the dielectric properties of bulk water, which has a dielectric constant of 88 at 0°C. A dielectric constant between 2 and 4 is usually used for the interiors of proteins. A larger effective dielectric constant, between 30 and 100, is considered appropriate for interactions involving unpaired charges, especially if the charges are solvent-accessible. The authors concluded that a value of eight should be appropriate for estimating the strength of short-range dipole-dipole interactions that occur near protein surfaces, such as those involved in ligand binding and enzymatic reactions.[11] In his review ‘Roles of electrostatic interactions in proteins’ (1996), Nakamura mentions a number of articles where the stabilizing influence of the helix dipole in short peptides is demonstrated. First of all there is the C-peptide, a thirteen residue fragment of ribonuclease A. It was found that Glu 2 and His 12, being negatively and positively charged respectively, are responsible for the stability of the peptide. Experiments were also conducted on the slightly larger fragment called the S-peptide. Various charges were introduced at the N-terminus, and the stability increased as the charge changed from +2 to -1. All experiments mentioned in the review agreed on the fact that helices do not necessarily have to be long to have a significant charge effect, even when the dielectric shielding effect was taken into account. The author concluded, based on the results of Aqvist and coworkers[1], that only the first one or two turns of an α-helix are likely to participate, and bring along a stabilization of at most about 2 kcal/mol, which is less then the estimate made by Hol.[8, 15] Following the discovery that the electrostatic field is drastically lowered by the field generated by the solvent, Sengupta et al. tried to calculate and quantify the net effective dipole moment using atomic-detail helix models and Poisson-Boltzmann continuum electrostatics calculations. Helices of varied lengths were used in different environments (vacuum, aqueous solution, lipid bilayers and protein interiors). The authors concluded that although the helix dipole is quite strong in vacuum, its strength is drastically reduced in aqueous solution due to the reaction field that is generated by the solvent. Further, while the net dipole moment increases with helix length in vacuum, the opposite is true for transmembrane helices, where an increasing length brings the termini closer to the shielding solvent. The authors state that the dipole moment of a helix can be used to determine electrostatic interactions at distances that are large compared to the dipole length and thus is likely inadequate for describing close range effects such as phosphate binding and antiparallel helix-motif stabilization. The authors postulate three interwoven rules-of thumb that follow from their results. The first one is that the dipole strength is determined by the position of the helix termini compared to the aqueous phase, whereas the total solvent-exposed area is less important. The second rule states that if both helix termini are solvent exposed, the net dipole moment will be small, almost equal to the situation where the helix is fully solvated. The third rule says that in cases where one helix terminus is solvent exposed and the other buried, asymmetric reaction field shielding can lead to a relatively high net dipole moment.[21] Lorieau and coworkers also reported an unexpected increase of pKa value for a highly conserved N-terminal glycine in the the influenza coat protein hemagglutinin HA2, which is known to mediate the fusion of viral and host-cell endosomal membranes. Two helices form a tight hairpin, stabilized by four interhelical hydrogen bonds. The authors argue that the reported pKa of 8.8 of the glycine is also due to interactions between the glycine and the dipole of the second helix. 3D NMR was applied, because protonations lead to a change in chemical shift. While the protonation state of the glycine was still in debate, the authors proposed the glycines amino group is in fact protonated. Normally, the pKa value of an N-terminal amino group of an alpha-helix in aqueous solution is depressed by about 0.5 units by the positive potential imposed by the helix dipole moment. Embedding of the peptide in the hydrophobic environment of the viral membranes would be expected to decrease its pKa value even further. However, a potential α-NH3+ interaction with one of the carboxyl groups of the peptide or with the phosphate of the lipid headgroup was proposed as a possible reason for the elevated pKa of this glycine.[12] Allosteric mechanisms In 2009, Preininger et al. demonstrated that electrostatics are also used to initiate nucleotide release in the mechanism of nucleotide exchange in G-proteins. Heterotrimeric G-proteins transmit signals from activated G protein-coupled receptors to downstream effectors through a guanine nucleotide signaling cycle. After the receptor binds its cognate G protein, the signal is transmitted from the receptor-binding site to the nucleotide binding pocket of the Gα-subunit. Numerous studies already suggested that the C-terminal α5 helix of Gα-subunits participates in Gα-receptor binding and that this binding process induced a rotational change in the same helix. Electron paramagnetic resonance (EPR) spectroscopy data indicated that the α5 helix rotation could be communicated to the nucleotide binding site both via global structural effects and through specific electrostatic effects directed at the base of the α5 helix, which moves closer to the bound guanine nucleotide upon rototranslation of the α5 helix. Using this knowledge and the protein structure of the Gα-subunit, the authors introduced two cysteines in the protein to lock the protein into its receptor-associated conformation via a disulfide bond. The 2.9 Å crystal structure of the protein in complex with GDPAlF4 proved the success of this disulfide bond. Because the roto-translation of the α5 helix would also be expected to move the positive dipole of the α5 helix toward the guanine ring of the bound nucleotide the authors constructed a second complex where a positive charge was introduced at the N-terminus, next to the TCAT-motif that is known to be involved in nucleotide binding. Since the negatively charged guanine ring may be attracted by the positively charged helix dipole, this construct was thought to lead to an increase in nucleotide exchange. Besides this front door exit for guanine, a back door mechanism may also exist. However, the authors cannot rule out one mechanism based on these experiments and additionally no helix dipole interactions are thought to be involved in the back door mechanism.[17] An example of a mechanism to switch between conformations where a helix dipole is involved is described by Neuwald. He analyzed both amino acid sequences and structures of several Ras-like GTPases. The on- and off-states of these GTPases is communicated through conformational changes in the switch I and II regions. Neuwald describes an additional regulatory structural element found in a subgroup of GTPases. At the end of the switch II helix an aromatic pocket is found with a positively charged residue inserted into it, facing the N-terminal side of the helix. This helix is pointed away from the GTPase core. Five residues that form the aromatic pocket are found to be highly conserved in these GTPases. The aromatic pocket is found in both the onand off-state. When this aromatic pocket is disrupted, the direction of the helix changes and comes along with a restructuring of the entire switch II Nterminal domain. This domain is known to sense the γ-phosphate of GTP and to harbor three previously noted Ras-like residues involved in GTP hydrolysis and nucleotide exchange. An association between the charge-dipole pocket and switch II restructuring is suggested by comparisons between typical Figure 7: The charge dipole pocket of GDP-bound Rab11a (PDB ID 1oiv, monomeric forms of Rab family indicates the characteristic atomic interactions of GTPases and an unusual homodimeric resolution 1.98 Å).The inset the charge dipole pocket.[16] form, Rab27, where the inter-switch region that connects switch I to switch II is exchanged between subunits. The switch II region forms a long α-helix that is directed away from the structural core of one subunit and toward the structural core of the other subunit. Other GTPases lack both the charge-dipole pocket configuration and the outward directed switch II helix. In addition, for GDP-bound Rab11a and Rab23, minor deviations from the charge-dipole pocket configuration are associated with shortening or lengthening of the outward-directed switch II helix, suggesting that this helix can extend and retract. Together, this indicates that the charge-dipole pocket configuration is closely associated, not with the on- or offstate, but rather with formation of the outward-oriented helix which, in turn, is associated with restructuring of the switch II N-terminal region and, as a result, with repositioning of co-conserved residues implicated in GTP hydrolysis and nucleotide exchange. Figure 7 shows the charge-dipole pocket and their interactions.[16] Somewhat similar to this protein is the bacterial transcription regulator YvoA, part of the GntR/HutC family which is investigated by Resch et al. At their N-terminus they contain a winged helix–turn–helix domain which is responsible for DNA binding. Both domains of the dimer are in close proximity in the DNA-bound conformation, binding in successive major grooves. The Cterminal domain of one unit of this homodimeric protein contains a chorismate lyase-type UTRA domain. Whereas this domain functions as an active site in other species, here it constitutes a binding pocket for various effector molecules.[18] This domain encompasses an allosteric mechanism that alters the affinity of the DNAbinding domains for DNA upon binding of an effector. Resch et al. introduced disulfide bonds to lock the regulator either in its DNA-bound or the effector-bound conformation in order to investigate the mechanism that triggers the transition between these conformations and its consequences for DNA binding affinity. The authors determined the structure with the transcription regulator bound to an effector, GlcNAc-6-P. They propose that the negatively charged phosphate group both stabilizes the helix dipole of one helix and induces the formation of another helix whose dipole also points toward the phosphate group. The new helix again induces a helix formation which generates internal symmetry in the effector binding domain and is proposed to be a key step in the propagation of the allosteric signal. The DNAbinding domains are then reoriented, rotating 122 degrees, Figure 8: The two conformations of YvoA[18] and hence unable to bind DNA. This mechanism, which the authors describe as a 'jumping jack'-like motion, is summarized in Figure 8.[18] Other helix types Another type of helix occurs in nature is the polyproline II helix, which is involved in biological processes such as signal transduction, transcription, immune response, and cell motility. Strands of collagen with the [Pro-Hyp-Gly]n repeat unit also adopt a PPII-like conformation. Kuemin et al. examinated the the influence of different terminal groups on the stability of this type of helix. Peptides with a PPII structure can switch to a PPI strucure, and the ease with which this occurred was used to measure differences in stability. An ideal PPII helix has dihedral angles φ and ψ of -75° and +145°, respectively, and all amide bonds are in trans conformations (ω = 180°). This results in a left-handed helix with every third residue stacked on top of each other in a lateral distance of 9.4 Å. Within this helix, all amide bonds of the peptide backbone are nearly perpendicular to the helix axis. The PPI helix is right-handed, and all amide bonds are in cis conformations with φ and ψ angles of -75° and +160°, respectively. These dihedral angles result in a more compact structure (helical pitch of 5.6 Å per turn, 3.3 proline residues per turn) with the amide bonds oriented nearly parallel to the helix axis. In contrast to the α-helix, neither the PPII nor the PPI helix is stabilized by intramolecular hydrogen bonds.[10] Four oligopeptides were used, similar in length but either carrying no terminal charges (at pH 7.0), a C-terminal negative charge, an N-terminal positive charge or both charges at the same time (AcN-[Pro]12- CONH2 (1), HN-[Pro]12-CONH2 (2), AcN-[Pro]12-CO2H (3), and HN[Pro]12-CO2H (4)). The PPII helix is the predominant conformation of oligoprolines in water, whereas the PPI helix is favoured in solvents such as n-PrOH. The amount of n-PrOH needed to switch the PPII helix to the PPI helix therefore reflects the ease of the conformational change between the PPII and the PPI form. The oligopeptides were studied in different mixtures of aqueous phosphate buffer (10 mM, pH 7.2) and n-PrOH by CD spectroscopy. The results revealed that capped termini stabilize the PPII helix, whereas charges have a destabilizing effect. This results from the fact that within the PPI helix, the individual dipoles are oriented almost linearly to the helix axis, and also being more compact. Moreover, the negative charge at the C-terminus has a more pronounced destabilizing effect as compared to a positive charge at the N-terminus. (95 vol % n-PrOH to switch whereas 85 vol % n-PrOH suffices for the C-terminal charge). The authors relate this difference to destabilizing effect by a repulsive interaction of the carboxylate with the oxygen of the neighbouring amide.[10] Supramolecular interactions Another example of interstrand stabilization is found in invertebrate collagen. Collagen is a supramolecular structure composed of three helices of individual strands. Shoulders et al. state that the amino acid composition of common collagen is composed of Gly-Xaa-Yaa repeats. Stable collagen triple helices form when (2S)-proline or Pro derivatives that prefer the Cγ-endo ring pucker are in the Xaa position and Pro derivatives that prefer the Cγ-exo ring pucker are in the Yaa position. Shoulders and Raines reported a form of collagen found in invertebrates that has a Cγ-exo–puckered Pro derivative ((2S,4R)-4-hydroxyproline, or Hyp), in the Xaa position. The triple helix is found to be hyperstable. After synthesizing analogous proteins lacking the possibility of obtaining additional stabilization through hydrophobic effects and increased hydrogen bonding[22]. A computational analysis by Improta, Berisio, and Vitagliano suggested that interstrand dipole–dipole interactions could more than compensate for the energetic penalty of triple-helix distortion caused by the incorporation of a Pro derivative with a Cγ-exo pucker in the Xaa position[9]. This proved to be the case, after synthesizing analogues where the presence of a favorable dipole–dipole interaction was the constant. The authors suggest that analogous strategies could be applicable in other contexts such as coiled-coils.[22] The supramolecular of collagen already made clear that the electrostatic effects of the helix dipole are not limited to intramolecular cases. Another case to illustrate this is the binding mechanism of centrin. Centrins are acidic Ca2+-binding proteins that are well conserved in the eukaryotic realm. Three isoforms of human centrin exist, with all isoforms containing two EF-hand motifs and target binding sites. Martinez-Sanz et al. examined the effects of a reversal of the traditional target sequence, and thus the reverse of the helix dipole, W1L4L8 on the dynamics and affinity of a single centrin isoform, HsCen2. A specific target sequence was chosen because an already determined protein structure featured this same sequence, albeit in the normal orientation. HsCen2 has the ability to bind any of the ∼25 repeats of this sequence, found in human Sfi1, with more or less affinity. As only the C-terminus was involved in the binding of the Sfi1 repeat, the structure was determined for the C-terminal domain only. The Trp in the target sequence is normally embedded in a hydrophobic cavity of HsCen2. With the reversed target sequence, this still proved to be the case, although the χ1 dihedral torsion angle switched from −80° to −160° . Another consequence was a ∼2 Å translation of the Sfi1 sequence towards a highly conserved glutamate sidechain, which is also conserved in calmodulin and troponin molecules for probably the same reasons. This resulted in a the lower enthalpy ΔΔH ≈ –7 kcal/mol and a lower free binding energy ΔΔG ≈ –2 kcal/mol compared to the normal orientation of Sfi1. This was measured using isothermal titration calorimetry. It is an aggregate effect, due to the a slight loss of hydrophobic interactions with Leu 8 and a inversion of the helix dipole that leads to a positive charge pointing at the conserved glutamate residue. The first of these probably exerts a negative effect on the calcium affinity of loop III of the centrin.[13] Conclusion Since the discovery of the helix dipole and its relevance to various biological structures and mechanisms, the subject has gained a serious foothold in the scientific literature. The exact amount of influence in reducing energetical costs, however, is still under debate. A number of researchers conclude, using both theoretical approaches and calculations and experiments, that only the termini are involved in this stabilization which according to them mainly originates from hydrogen bonding. Nevertheless, its overall biological relevance is widely acknowledged References 1. Aqvist, J., Luecke, H., Quiocho, F. A., Warshel, A. (1991). Dipoles localized at helix termini of proteins stabilize charges, Proc. Natl. Acad. Sci. USA 88 2026-2030. 2. Born, M. (1920). Volumen und Hydratationswärme der Ionen. Zeitschrift für Physik A Hadrons and Nuclei 1 45-48. 3. Chatelain, F.C., Alagem, N., Xu, Q., Pancaroglu, R., Reuveny, E., Minor Jr., D.L. (2005). The Pore Helix Dipole Has a Minor Role in Inward Rectifier Channel Function. Neuron 47(6) 833–843. 4. Doyle, D.A., Cabral, J.M., Pfuetzner, R.A., Kuo, A., Gulbis, J.M., Cohen, S.L., Chait, B.T., MacKinnon, R. (1998). The Structure of the Potassium Channel: Molecular Basis of K+ Conduction and Selectivity. Science 280 69-77. 5. Dutzler, R., Campbell, E.B., Cadene, M., Chait, B.T., MacKinnon, R. (2002). X-ray structure of a ClC chloride channel at 3.0 Å reveals the molecular basis of anion selectivity. Nature 415 287-294. 6. Faraldo-Gomez J.D. and Roux B. Electrostatics of Ion Stabilization in a ClC Chloride Channel Homologue from Escherichia coli. J. Mol. Biol. 339 981–1000. 7. Fu, D., Libson, A., Miercke, L.J.W., Weitzman, C., Nollert, P., Krucinski, J., Stroud, R.M. (2000). Structure of a glycerol-conducting channel and the basis for its selectivity. Science 290 481±486. 8. Hol, W.J G. (1985). The role of the α-helix dipole in protein function and structure, Prog. Biophys. molec. Biol. 45 149-195. 9. Improta, R., Berisio, R., and Vitagliano, L. (2008). Contribution of dipole-dipole interactions to the stability of the collagen triple helix. Protein Science 955-961 10. Kuemin, M., Schweizer, S., Ochsenfeld, C., Wennemers, H, (2009). Effects of Terminal Functional Groups on the Stability of the Polyproline II Structure: A Combined Experimental and Theoretical Study. J. Am. Chem. Soc. 131 15474–15482. 11. Lockhart, D.J., Kim, P.S. (1992). Internal stark effect measurement of the electric field at the amino terminus of an alpha helix. Science 257 947–951. 12. Lorieau, J.L., Louis, J.M., Bax, A. (2011). Helical Hairpin Structure of Influenza Hemagglutinin Fusion Peptide Stabilized by Charge-Dipole Interactions between the N-Terminal Amino Group and the Second Helix. J. Am. Chem. Soc. 133 2824–2827. 13. Martinez-Sanz, J.,Kateb, F., Assairi, L., Blouquit, Y., Bodenhausen, G., Abergel, D., Mouawad, L., Craescu, C.T.. (2010). Structure, Dynamics and Thermodynamics of the Human Centrin 2/hSfi1 Complex. J. Mol. Biol. 395 191–204. 14. Murata, K.,Mitsuoka, K., Hirai, T., Walz, T., Agrek, P., Heymann, J.B., Engel, A., Fujiyoshi, Y. (2000). Structural determinants of water permeation through aquaporin-1. Nature 407 599±605. 15. Nakamura H. (1996) Roles of electrostatic interaction in proteins. Q Rev Biophys. 29 1–90. 16. Neuwald A. (2009). The Charge-dipole Pocket: A Defining Feature of Signaling Pathway GTPase On/Off Switches. J. Mol. Biol. 390 142–153. 17. Preininger, A.M.,Funk, M.A., Oldham, W.M., Meier, S.M., Johnston, C.A., Adhikary, S., Kimple, A.J., Siderovski, D.P., Hamm, H.E., Iverson, T.M. (2009). Helix Dipole Movement and Conformational Variability Contribute to Allosteric GDP Release in GRi Subunits. Biochemistry 48 2631. 18. Resch M., Schiltz, E., Titgemeyer, F., Muller Y.A. (2010). Insight into the induction mechanism of the GntR/ HutC bacterial transcription regulator YvoA. Nucleic Acids Research 38 2485–2497. 19. Roux B., MacKinnon R. (1999). The cavity and pore helices in the KcsA K+ channel: electrostatic stabilization of monovalent cations. Science 285 100–102. 20. Sali D., Bycroft, M., Fersht, A.R. (1988) Stabilization of protein structure by interaction of alpha-helix dipole with a charged side chain. Nature 20 740-3. 21. Sengupta D., Behera, R.N., Smith, J.C., Ullmann, G.M. (2005). The α Helix Dipole: Screened Out? Structure 13 849–855. 22. Shoulders, M.D., Raines, R.T. (2011). Interstrand Dipole–Dipole Interactions Can Stabilize the Collagen Triple Helix. J Biol Chem. 286(26) 22905-12. 23. Zhou, Y., Morais-Cabral, J.H., Kaufman, A., MacKinnon, R.. (2001). Chemistry of ion coordination and hydration revealed by a K+channel±Fab complex at 2.0 Å resolution. Nature 414 43±48.